phil. trans. r. soc. lond. 2002, 357, 1481

Upload: enrique-pradias

Post on 06-Apr-2018

223 views

Category:

Documents


0 download

TRANSCRIPT

  • 8/3/2019 Phil. Trans. R. Soc. Lond. 2002, 357, 1481

    1/18

    Published online 3 October 2002

    The design and synthesis of artificial photosynthetic

    antennas, reaction centres and membranes

    T. A. Moore*, A. L. Moore and D. Gust

    Department of Chemistry and Biochemistry and Centre for the Study of Early Events in Photosynthesis,

    Arizona State University, Tempe, AZ 85287-1604, USA

    Artificial antenna systems and reaction centres synthesized in our laboratory are used to illustrate that

    structural and thermodynamic factors controlling energy and electron transfer in these constructs can

    be modified to optimize performance. Artificial reaction centres have been incorporated into liposomal

    membranes where they convert light energy to vectorial redox potential. This redox potential drives a

    Mitchellian, quinone-based, proton-transporting redox loop that generates a H of ca. 4.4 kcal mol1

    comprising pH ca. 2.1 and ca. 70 mV. In liposomes containing CF0F1ATP synthase, this system

    drives ATP synthesis against an ATP chemical potential similar to that observed in natural systems.

    Keywords: artificial photosynthesis; artificial antenna; artificial reaction centres

    1. INTRODUCTION

    Natural photosynthesis has inspired chemists to mimic

    aspects of this successful energy-transducing process in

    the laboratory. One of many approaches to the assembly

    of an artificial reaction centre is to use synthetic pigments,

    electron donors and electron acceptors that are related to

    the natural pigments (chlorophylls, carotenoids and

    quinones), but to employ covalent bonds in place of the

    proteins that function as scaffolds, organizing the pig-

    ments in the natural systems. Like the natural proteinmatrix, the covalent bonds of the artificial antennae and

    reaction centres control the separations, angular relation-

    ships and electronic coupling among the various active

    moieties, and thus control the rates and yields of energy

    and electron-transfer processes (Wasielewski 1992; Gust

    et al. 1993a,b, 1999, 2000, 2001a,b; Gust & Moore 2000).

    In 3, we will discuss biomimicry of the various functions

    of carotenoids in photosynthesis. Later, in 5, the prep-

    aration and function of artificial reaction centres will be

    described. Finally, in 6 we will review progress in the

    construction of artificial photosynthetic membranes.

    2. CAROTENOIDS IN NATURAL PHOTOSYNTHESIS

    (a) Carotenoid pigments as photosynthetic

    antennae

    As accessory light-harvesting pigments, carotenoids are

    found in the chlorophyll-binding antenna proteins of

    photosynthetic organisms where they absorb light in the

    bluegreen spectral region and transfer energy to nearby

    chlorophylls.

    Carhv

    1Car, (2.1)

    * Author for correspondence ([email protected]).

    One contribution of 21 to a Discussion Meeting Issue Photosystem II:

    molecular structure and function.

    Phil. Trans. R. Soc. Lond. B (2002) 357, 14811498 1481 2002 The Royal SocietyDOI 10.1098/rstb.2002.1147

    1Car ChlCar 1Chl. (2.2)

    Light energy collected in this way is funnelled to special

    structures in the membranes, known as reaction centres,

    where energy conversion to electrochemical potential

    takes place. The unique photophysics of carotenoids (vide

    infra) places severe constraints on aspects of the structure

    of carotenoidchlorophyll-binding proteins that control

    the interactions of the pigments. As described below,many of these structural constraints have been identified

    in model systems by making systematic changes in the

    linkages connecting carotenoid pigments to tetrapyrroles

    and measuring the effect on energy-transfer rates and

    yields. We begin with a brief review of carotenoid photo-

    physics to set the stage for interpreting energy-transfer

    experiments on model antenna systems.

    (b) Carotenoid photophysics

    This discussion applies to the genre of carotenoid pig-

    ments having nine or more conjugated double bonds,

    including oxy- and hydroxy-substituted compounds,

    although there are certainly small differences between pig-ments. Carotenoid absorption of light in the 400550 nm

    spectral region gives rise to their intense yelloworange

    colour, a consequence of high extinction coefficients (ca.

    105 M1 cm1) in this region. This strongly electric dipole

    allowed transition is from the ground state (S0) to the

    second excited singlet state (S2) of the carotenoid. As

    expected for an upper excited singlet state, the S2 species

    has a very short lifetime of 200 fs or less (Truscott 1990;

    Shreve et al. 1991, 1992; Kandori et al. 1994; Macpher-

    son & Gillbro 1998). Because of the short lifetime, the

    quantum yield of fluorescence emission from this state is

    very low (less than 104) in spite of the strongly allowed

    nature of the transition and its correspondingly large radi-ative rate constant (ca. 109 s1). Surprisingly, this state

    is important as an energy donor in photosynthesis

    (Trautman et al. 1990; Ricci et al. 1996).

  • 8/3/2019 Phil. Trans. R. Soc. Lond. 2002, 357, 1481

    2/18

    1482 T. A. Moore and others Artificial antenna systems and reaction centres

    The transition from S0 to the lowest excited singlet state

    of carotenoids, S1, is electric-dipole forbidden (Kohler

    1991) and is not observed in the usual single-photon

    absorption experiment. It is readily populated by relax-

    ation from S2. The lifetime of the S1 state, as measured by

    transient absorption techniques, is ca. 1040 ps for most

    carotenoids (Wasielewski & Kispert 1986; Trautman et al.

    1990). The state decays essentially exclusively via internalconversion. Fluorescence from S1 is virtually undetectable

    (quantum yield less than 104) and intersystem crossing

    to the triplet state is not observed. The energy of the S1state is uncertain for most of the carotenoids involved in

    photosynthesis because the forbidden nature of the

    S0 S1 transition makes it difficult to detect.

    Although carotenoid triplet states are not formed in

    appreciable amounts by the usual intersystem crossing

    pathway, they can be produced via triplettriplet energy

    transfer from other species or by radical-pair recombi-

    nation processes in artificial reaction centres (Gust et al.

    1992; Liddell et al. 1997; Carbonera et al. 1998). Thus,

    their transient absorption characteristics are well known(Mathis & Kleo 1973; Bensasson et al. 1976). Typically,

    they have strong absorption maxima at about 540 nm. In

    most organic solvents in the absence of oxygen, carotenoid

    triplet states have lifetimes of 510 s. Phosphorescence

    from these triplets has not been reported, and therefore

    the energies of the carotenoid triplet states have not been

    measured by direct optical means. Energy-transfer studies

    have demonstrated that the energy of the lowest-lying car-

    otenoid triplet is below that of singlet oxygen at ca. 1.0 eV.

    Results from our laboratory were obtained using a hybrid

    spectroscopiccalorimetric technique (photoacoustic

    spectroscopy) that has located the lowest-lying triplet state

    of a representative carotenoid (similar to the carotenoidmoiety of structure 10 (see Appendix A) except containing

    one more conjugated double bond) at 0.65 0.04 eV

    (Lewis et al. 1994). Thus, the expected origin of the phos-

    phorescence would be ca. 1900 nm.

    (c) Photoprotection

    In the photosynthetic process, carotenoids have func-

    tions that are ancillary to the basic energy-conversion reac-

    tions and are crucial to the protection of photosynthetic

    membranes from photochemical damage (Cohen-Bazire &

    Stanier 1958; Sieferman-Harms 1987; Frank & Cogdell

    1996). Under certain conditions, most photosynthetic

    organisms unavoidably generate chlorophyll triplet excitedstates when exposed to light. Often, these are produced in

    the reaction centre via recombination of charge-separated

    states. Chlorophyll triplet states are excellent sensitizers

    for formation of singlet oxygen from the oxygen ground-

    state triplet as in the following equation:

    3Chl 3O2Chl 1O2. (2.3)

    Singlet oxygen, with an energy of ca. 1 eV above the

    ground state, is a highly reactive species that can react

    with and destroy lipid bilayer membranes and other vital

    components of cells. Clearly, singlet oxygen production in

    photosynthetic organisms is potentially injurious (Krinsky

    1979; Cogdell & Frank 1993). Such organisms havedeveloped dual defence mechanisms based upon the low

    energy of the carotenoid triplet state mentioned above.

    Carotenes can quench chlorophyll triplet states at a rate

    Phil. Trans. R. Soc. Lond. B (2002)

    that precludes singlet oxygen generation via a triplet

    triplet energy-transfer process (equation (2.4)). Addition-

    ally, carotenoids can quench singlet oxygen itself through

    energy transfer (equation (2.5)) or chemical reaction

    (Foote et al. 1970).

    3Chl CarChl 3Car, (2.4)

    1O2 Car3O2

    3Car. (2.5)

    Because the carotenoid triplet state is much lower in

    energy than singlet oxygen, it returns harmlessly to the

    ground state with the liberation of heat.

    From this overview, it is readily appreciated that carot-

    enoids were selected by the evolving photosynthetic appar-

    atus because of their unique photophysical properties.

    Their strong absorption is closely matched to the

    maximum of the solar flux, although their triplet levels are

    well below the energy of singlet oxygen.

    (d ) Regulation of photosynthesis by carotenoids

    It has been found that carotenoid polyenes are involved

    in a mechanism whereby some photosynthetic organisms

    dissipate excess chlorophyll singlet excitation energy

    (Yamamoto 1979; Demmig-Adams 1992; Yamamoto &

    Bassi 1995; Horton et al. 1996). This process is sometimes

    referred to as non-photochemical quenching because it is

    measured by in vivo fluorescence spectroscopy of plant

    leaves. It comes into play at high light levels and evidently

    helps to protect the organism from light-induced damage.

    Non-photochemical quenching is signalled by the conver-

    sion of violaxanthin to zeaxanthin, but the details of the

    quenching process are currently not understood.

    3. CAROTENOIDS IN BIOMIMETIC SYSTEMS

    (a) Mimicry of the antenna function of carotenoids

    As indicated in equation (2.2), in order to act as

    antennas for gathering light, carotenoid pigments must

    transfer singlet excitation energy to nearby chlorophyll

    molecules. Singlet energy transfer from carotenoid to

    chlorophyll can occur by at least two different pathways.

    At one limit, following absorption of light and internal

    conversion from S2, the S1 level of the carotenoid is popu-

    lated and then transfers energy to the Qy (S1) of the

    chlorophyll. At the other limit, energy transfer from the

    S2 level of the carotenoid to the Q x (S2) of the chlorophyllcompetes with internal conversion in the carotenoid. The

    short lifetime of the S2 state of the carotenoid does not a

    prioriexclude the second possibility. Indeed, detailed spec-

    troscopic studies on the subpicosecond time-scale of the

    pigment protein complexes that make up the light-

    harvesting antennas of photosynthetic organisms indicate

    that, in some cases, energy transfer from the carotenoid

    to the chlorophyll or bacteriochlorophyll may occur from

    the carotenoid S2 level rather that from the lower-energy

    S1 excited state, as predicted by Kashas rule (Trautman

    et al. 1990; Shreve et al. 1991; Koyama et al. 1996; Ricci

    et al. 1996; Krueger et al. 1998). The two pathways are

    shown schematically in figure 1.However, the complexity of the natural systems con-

    taining multiple chromophores, coupled with the difficulty

    of the measurements on such a short time-scale, make the

  • 8/3/2019 Phil. Trans. R. Soc. Lond. 2002, 357, 1481

    3/18

    Artificial antenna systems and reaction centres T. A. Moore and others 1483

    energy(eV)

    1.0

    0

    2.0

    carotenoid tetrapyrrole

    hv

    Qx

    Qy

    S0S0(11Ag_

    )

    S1(21Ag_)

    S2(11Bu

    +)

    Figure 1. Energy-level diagram showing both possible paths

    for energy transfer from the excited singlet states of a

    carotenoid to the S1 or S2 levels of a tetrapyrrole pigment.

    pathway assignments uncertain. Structures 1 and 4 (see

    Appendix A) are examples of simple, well-defined molecu-

    lar dyads that were designed to help in the understanding

    of the carotenoid to chlorophyll singletsinglet energy-

    transfer process in natural systems.Transient absorption and fluorescence experiments on

    structures 1 and 4 and the model pigments, structures 2,

    3, 5 and 6 (see Appendix A), yielded convincing results

    about the energy-transfer pathway (Macpherson et al.

    2002). In structure 1, the lifetime of S2 was measured by

    fluorescence upconversion to be 45 fs, whereas the S2 life-

    time of model carotenoid 3 is 160 fs. The S 1 lifetime was

    8 ps in both structures 1 and 3. Moreover, the S 1 rise time

    for tetrapyrrole (moiety of dyad 1) was found to be 62 fs

    by fluorescence upconversion. Therefore, only the carot-

    enoid S2 state was quenched by the attached tetrapyrrole.

    The observation that the time constant associated with the

    decay of the energy donor (carotenoid S2) is consistentwith that associated with the energy arrival at the acceptor

    (tetrapyrrole S1) solidly demonstrates the S2 energy-

    transfer pathway. (The difference between carotenoid S2decay at 45 fs and tetrapyrrole S1 rise at 62 fs is due to

    the delay associated with relaxation to the fluorescent state

    in the tetrapyrrole.) Quantitatively, quenching of the caro-

    tenoid S2 state from 160 fs to 45 fs by energy transfer

    results in an efficiency of 70%, which matches that meas-

    ured by steady-state fluorescence excitation spectroscopy.

    Studies of dyad 4 illustrate that both pathways can be

    active. In initial experiments on dyad 4, the attached tetra-

    pyrrole was found to quench the carotenoid S2 state from

    95 fs in model 6 to 28 fs, and the carotenoid S1 level from12 to 9 ps. The rise of the S1 level of the tetrapyrrole of

    structure 4, as measured by fluorescence upconversion,

    required a major exponential component (74%) of 41 fs

    Phil. Trans. R. Soc. Lond. B (2002)

    and a minor component (26%) of 4 ps. While the match

    between the 9 ps decay of the carotenoid S1 and the 4 ps

    rise component of the tetrapyrrole S1 is only qualitative,

    these preliminary experiments do provide evidence that

    both states can be energy donors. Taken together, the car-

    otenoid S2 and S1 kinetic data give an overall quantum

    yield of energy transfer of ca. 80%, in qualitative agree-

    ment with that observed by steady-state fluorescence-excitation spectroscopy.

    Why is the yield of the S1 pathway different in structures

    1 and 4? There are undoubtedly subtle differences in the

    electronic coupling between the chromophores (vide

    supra) and the energy-transfer energetics are different. The

    S1 level of the tetrapyrrole in stucture 4 is about 270 cm1

    higher than that of structure 1. Considering this small

    energy difference, the fact that the S1 carotenoid band is

    expected to be rather broad (Frank et al. 1996) and that

    the S1 levels of these carotenoids have not been measured

    but should be similar, the role of thermodynamics in con-

    trolling these energy-transfer processes remains speculat-

    ive.

    (b) Carotenoid antenna for non-photosynthetic

    pigments

    One of the clearest examples of the increased absorption

    cross-section for a photochemical process provided by

    carotenoid pigments is that observed in carotenoid

    buckminsterfullerene dyad, structure 7 (see Appendix A)

    (Imahori et al. 1995). The carotenoid absorption spec-

    trum is distinct and much stronger than that of the under-

    lying C60 bands. Upon excitation of the carotenoid moiety

    of stucture 7 with a 150 fs pulse of 600 nm laser light, the

    carotenoid radical cation forms in ca. 800 fs via photo-

    induced electron transfer to the fullerene to yieldCC60. The charge-separated species has a lifetime of

    ca. 500 ps and decays partially to the carotenoid triplet.

    Excitation of the carotenoid pigment provides two routes

    to the charge-separated species: energy transfer from the

    carotenoid to the C60 followed by photoinduced electron

    transfer to the C60 S1 and direct photoinduced electron

    transfer from the carotenoid S1 level. Although the contri-

    butions of the two pathways leading to CC60 could not

    be separated, the overall quantum yield for the formation

    of CC60 is unity. These results illustrate one way that

    primary photochemistry with a high yield can occur upon

    excitation of a carotenoid pigment, and keep the door

    open to the suggestion that carotenoids can serve as bluelight photoreceptors (Quinones et al. 1996).

    (c) Mimicry of the regulation of photosynthesis:

    non-photochemical quenching

    Regarding the role of carotenoid pigments in non-

    photochemical quenching of photosynthesis mentioned in

    1, it is interesting to compare structures 1 and 4. Both

    demonstrate a high level of antenna function as measured

    by energy-transfer efficiencies of 70 to 80%, which are

    similar to those observed in many natural systems.

    Regarding the tetrapyrrole singlet energies, the S1 level of

    structure 1 is slightly lower than that of structure 4.

    Although the carotenoid moieties in structures 1 and 4 aredifferent and their S1 levels are not known, there is no

    evidence from the observed absorption into S2 that there

    are significant differences in their singlet energy levels.

  • 8/3/2019 Phil. Trans. R. Soc. Lond. 2002, 357, 1481

    4/18

    1484 T. A. Moore and others Artificial antenna systems and reaction centres

    Interestingly, in structure 4 the tetrapyrrole fluorescence

    is quenched (by the attached carotenoid) ca. 10-fold over

    that of structure 1. This is consistent with a simple

    description of the tetrapyrrole energetics in that energy-

    transfer quenching from the tetrapyrrole S1 to the carot-

    enoid would be more likely in structure 4 because the

    tetrapyrrole S1 level is above that of the tetrapyrrole of

    structure 1, and is therefore more likely to be above thatof the carotenoid S1. However, this interpretation is incon-

    sistent with the carotenoid S1 state acting as an energy

    donor in structure 4 and not in structure 1. If tetrapyrrole

    to carotenoid energy transfer in structure 4 is thermodyn-

    amically downhill, then carotenoid to tetrapyrrole energy

    transfer must be uphill and is more uphill in structure 4

    than in structure 1. Although we do not have an expla-

    nation for the large quenching effect and enigmatic

    energy-transfer behaviour, structures 1 and 4 do illustrate

    that subtle changes in parameters such as the coupling

    between the pigments, their redox levels and their spectro-

    scopic energies can modulate and direct the energy flow

    between carotenoids and tetrapyrroles.What is the role of energy or electron transfer in the

    quenching of tetrapyrrole fluorescence by carotenoids?

    In other model studies, the redox levels of a porphyrin

    carotenoid dyad have been shown to influence the

    quenching mechanism to the extent that electron transfer

    from the carotenoid to the excited porphyrin was shown

    to occur (Hermant et al. 1993). However, in a series of

    carotenoidporphyrin dyads in which the number of con-

    jugated carboncarbon double bonds in the carotenoid

    moiety was systematically increased from 7 to 11, quench-

    ing could not be assigned uniquely to either electron- or

    energy-transfer processes (Cardoso et al. 1996). Unfortu-

    nately, these model studies have not produced a definitiveanswer to the electron-/energy-transfer question.

    (d ) Mechanistic considerations of energy transfer

    Regarding the role of carotenoids as antenna pigments,

    energy transfer from them to chlorophylls can be brought

    about by a variety of interactions. If the distance between

    the chromophores is larger than their transition dipole

    lengths, if the transition dipoles have sufficient strength

    and the correct orientation, and if the donor state has suf-

    ficient lifetime (usually expressed as fluorescence quantum

    yield), energy transfer may be described by the Forster

    mechanism (Forster 1948). As the chromophores are

    brought closer together, higher-order electrostatic termsmust be included in the electronic coupling matrix

    elements. The inclusion of these terms can relax the need

    for strong, dipole-allowed transitions and relatively long-

    lived states in the donor and can account for the carot-

    enoid S1 and S2 levels acting as donor states (Krueger et

    al. 1998). If the chromophores are brought into van der

    Waals contact so that there is orbital overlap between the

    donor and the acceptor, the Dexter electron-exchange

    mechanism becomes important.

    The Dexter mechanism does not depend on dipole or

    mutipole strengths and is the accepted mechanism for

    triplettriplet energy transfer. How fast can Dexter-

    mediated energy transfer be and to what extent do elec-tron exchange-based electronic coupling matrix elements

    contribute to singlet energy transfer in these systems? In

    dyads 1 and 2 and several other carotenoid-containing

    Phil. Trans. R. Soc. Lond. B (2002)

    multichromophoric molecules, energy transfer from the

    triplet tetrapyrrole to populate the carotenoid triplet state

    is remarkably fast. Due to kinetic constraints inherent to

    these dyads, the actual rate cannot be determined

    (Macpherson et al. 2002). In these model systems, the rise

    time of the carotenoid triplet energy acceptor is the same

    as the singlet lifetime of the tetrapyrrole triplet energy

    donor. This indicates that intersystem crossing in thedonor is the rate-limiting step in the flow of energy from

    the singlet of the donor to the triplet of the acceptor. In

    other words, the true rise time of the triplet carotenoid is

    much faster than intersystem crossing in the tetrapyrrole,

    so that the carotenoid triplet is observed to rise with the

    singlet lifetime of the tetrapyrrole.

    The final answer to the second part of the question

    awaits detailed quantum chemical calculations. In the

    meantime, it has been addressed experimentally in one

    study of three isomeric carotenoporphyrins in which singlet

    singlet energy-transfer efficiency (albeit at much lower

    efficiency than observed in dyads 1 and 2) and triplet

    triplet energy-transfer rates were found to depend in thesame way on the electronic structure of the interchromo-

    phore linkage group (Gust et al. 1992). Because the triplet

    energy-transfer process is mediated by the linkage group

    via the Dexter mechanism, the simplest interpretation is

    that in those dyads the singlet energy transfer is also elec-

    tron-exchange mediated.

    4. EVOLUTION OF CAROTENOID

    ANTENNA/PHOTOPROTECTIVE FUNCTION

    IN PHOTOSYNTHESIS

    One may assume that carotenoids were selected by the

    evolving photosynthetic apparatus because of their uniquephotophysical properties. Three properties can be ident-

    ified that stand out in this regard: (i) a highly allowed

    absorption band in the 450 to 500 nm range, where the

    solar flux is high for efficient light harvesting; (ii) an S1energy level that is highly forbidden (vide supra); and (iii)

    a lowest-excited triplet state that is well below the energy

    of singlet oxygen (1g, 0.98 eV) so that equation (2.5) is

    irreversible. Notwithstanding the fact that the highly

    allowed absorption band near the solar irradiance

    maximum accounts for between 25 and 50% of the sun-

    light that drives photosynthesis on the earth, the crucial

    role of carotenoids in photosynthesis today is undoubtedly

    to protect photosynthetic organisms from the deleteriouseffects of singlet oxygen.

    It is interesting to speculate about the evolutionary ori-

    gin of photoprotection and the link between photoprotec-

    tion and the highly forbidden carotenoid S1 level (Moore

    et al. 1990, 1994). From the earliest photosynthetic organ-

    isms until the evolution of oxygen production (2.5ca. 3.5

    billion years ago), atmospheric oxygen levels were

    extremely low; there would have been no need for protec-

    tion from singlet oxygen under these conditions. However,

    it is almost certain that photosynthetic organisms evolving

    during this time used carotenoids as light-harvesting pig-

    ments. In order to do so, chlorophyll and carotenoid bind-

    ing proteins evolved in which singlet energy transfer fromcarotenoids to chlorophylls was efficient. Because of the

    forbidden nature of the carotenoid S1 level, efficient

    energy transfer required extensive electronic coupling

  • 8/3/2019 Phil. Trans. R. Soc. Lond. 2002, 357, 1481

    5/18

  • 8/3/2019 Phil. Trans. R. Soc. Lond. 2002, 357, 1481

    6/18

    1486 T. A. Moore and others Artificial antenna systems and reaction centres

    absorption and emission spectra. The fullerene absorbs

    throughout the visible spectrum with a threshold at ca.

    710 nm; its extinction coefficients in the visible region are

    much lower than those of the porphyrin. The absorption

    spectra are essentially linear combinations of those of the

    component chromophores, which means that the spectra

    show no evidence of strong electronic interaction of the

    chromophores. Cyclic voltammetric studies on structure9 (see Appendix A) and model compounds allow esti-

    mation of the energy of PC60 as 1.58 eV above the

    ground state in polar solvents.

    Analysis of time-resolved fluorescence and absorption

    data for the dyads and appropriate model compounds

    allows calculation of rate constants for various photo-

    chemical processes. In toluene, 1PC60 of structure 9

    decays by singletsinglet energy transfer (step 1 in figure

    2; k1 = 4.5 1010 s1) to yield P1C60, which undergoes

    intersystem crossing to the triplet state. No photoinduced

    electron transfer is observed. In polar solvents such as

    benzonitrile, 1PC60 decays by a combination of photoind-

    uced electron transfer by step 2 (k2 = 1.8 1011

    s1

    ) togive PC60 and singlet energy transfer by step 1

    (k1 = 7.1 1010 s1). The P1C60 formed in this manner

    or by direct absorption of light decays by photoinduced

    electron transfer step 3 to also yield the PC60 state

    (k3 = 1.3 1010 s1). The charge-separated state is easily

    identified using transient absorption spectroscopy by the

    absorption of the porphyrin radical cation in the 600

    700 nm region and the fullerene radical anion at ca.

    1000 nm. The overall quantum yield of charge separation

    via both pathways is 0.99. The PC60 state decays

    by charge recombination to the ground state with

    k4 = 3.4 109 s1.

    In general, the studies of photoinduced electron transferin porphyrinfullerene dyads have shown that charge-

    separated states are produced in high yield, and have long

    lifetimes relative to those of similar porphyrinquinone

    systems (Liddell et al. 1994; Kuciauskas et al. 1996). For

    example, a comparison of the photochemistry of a porphyrin

    fullerene dyad with that of a structurally related porphyrin

    quinone dyad shows that in tetrahydrofuran solution,

    photoinduced electron transfer in the C60-dyad occurs

    with a rate constant ca. six times larger than that in the

    quinone-based dyad, even though the latter one has a driv-

    ing force that is larger by 0.28 eV (Imahori et al. 1996).

    But the most remarkable difference is in the charge recom-

    bination reaction. In the quinone-based dyad, this reac-tion occurs with a rate constant of ca. 25 times greater

    than in the C60-based dyad. The difference in rate con-

    stants has been ascribed to a significantly smaller reorgani-

    zation energy for the large, diffuse fullerene radical anion

    than for the quinone radical anion, where the negative

    charge is mostly concentrated in the region of the oxygen

    atoms (Imahori et al. 1996; Guldi 2000; Kuciauskas et al.

    2000). Because of the small , the maximum of the Mar-

    cus curve occurs at a numerically smaller G, shifting

    the more exergonic recombination reaction well into the

    Marcus inverted region and slowing recombination.

    (c) Triad-based artificial reaction centresDyad systems can generate charge separation with a

    quantum yield of essentially unity while preserving a large

    part of the photon energy as intramolecular redox poten-

    Phil. Trans. R. Soc. Lond. B (2002)

    tial. However, rapid charge recombination to the ground

    state, which usually occurs on the subnanosecond time-

    scale, releases the stored energy as heat; it is difficult to

    devise intermolecular energy-conserving processes that

    can compete with this degradation.

    In the early 1980s, we designed an artificial photosyn-

    thetic reaction centre that overcomes this problem by

    using a multistep electron-transfer strategy such as thatfound in natural reaction centres (Moore et al. 1984).

    Recently, we have designed a series of triad arti ficial reac-

    tion centres based on the dyads described above. These

    constructs illustrate that careful consideration of the ther-

    modynamic and electronic factors controlling electron

    transfer can result in artificial reaction centres that gener-

    ate charge-separated states that rival those from natural

    reaction centres in quantum yield, energetics and lifetime.

    (d ) Photochemistry of fullerene-based triad

    artificial reaction centres

    As described above, photoinduced electron-transfer

    studies in porphyrinfullerene dyads demonstrated thatcharge-separated states can be produced in high yield and

    that they have long lifetimes relative to quinone-containing

    dyads. These properties of fullerene systems indicate that

    they might be ideal components of more complex systems,

    where slow charge recombination would favour evolution

    of a charge-separated state into a longer-lived species via

    multistep electron transfers similar to the processes

    observed in the reaction centres of photosynthetic organ-

    isms.

    In 1997, we reported a next step in the evolution of

    such molecules in the form of carotene (C) porphyrin (P)

    fullerene (C60) triad 10 (Gust et al. 1997; Liddell et al.

    1997; Kuciauskas et al. 2000). Other triads based on C60have subsequently been reported (Luo et al. 2000). The

    photochemistry of this triad may be discussed with refer-

    ence to figure 3, which shows the relevant high-energy

    states and decay pathways. Time-resolved spectroscopic

    studies have shown that at ambient temperatures in 2-

    methyltetrahydrofuran solution, excitation of the porphy-

    rin moiety yields C1PC60, which decays in 10 ps by pho-

    toinduced electron transfer to the fullerene to generate

    CPC60 (step 3 in figure 3; k3 = 1.0 1011 s1). Also,

    the fullerene excited singlet state accepts an electron from

    the porphyrin to yield CPC60 (k2 = 3.1 1010 s1).

    The overall quantum yield of CPC60 by these two

    pathways is essentially unity. Secondary electron transferfrom the carotenoid (step 4) competes with charge recom-

    bination by step 7 to give the final CPC60 charge-

    separated state. The rise time of this state is 80 ps and the

    overall yield is ca. 0.14, as determined by the comparative

    method. The CPC60 state decays in 170 ns exclus-

    ively by charge recombination to produce the carotenoid

    triplet state (step 9, t = 0.13). Decay of3CPC60 occurs

    in 4.9 s (k15 = 2.0 105 s1). Even in a glass at 77 K,

    CPC60 is formed from C1PC60 with a quantum

    yield of ca. 0.10 and again decays to give 3CPC60. The

    fullerene triplet, formed from CP1C60 (step 10) through

    normal intersystem crossing, is also observed at 77 K.

    Importantly, charge recombination to form a tripletspecies in triad 10 provides a spectroscopic label to follow

    the flow of triplet energy and the spin dynamics in model

    photosynthetic reaction centres. The spin-polarized EPR

  • 8/3/2019 Phil. Trans. R. Soc. Lond. 2002, 357, 1481

    7/18

    Artificial antenna systems and reaction centres T. A. Moore and others 1487

    ...

    0

    .

    Figure 3. Energy-level diagram showing relevant high-energy states and energy and electron-transfer decay pathways for triad

    10. Triads 11 and 12 differ in the energies of certain charge-separated states as fully described in 5d. The singlet radical-pairstate 1(CPC60) is shown along with the three triplet radical-pair states

    3(CPC60). The splittings of the triplet

    radical-pair state energies are not to scale.

    signal of the final charge-separated state, CPC60,

    was observed and simulated (Hasharoni et al. 1990). The

    fact that both electron-transfer steps in structure 10 occur

    even in a glassy matrix at 77 K has made it possible to

    observe some unusual and biologically relevant triplet

    triplet energy-transfer phenomena in the triads (Gust et al.

    1997). In triad 10, triplet energy is transferred from the

    fullerene moiety of CP3

    C60 (generated by intersystemcrossing) to the carotenoid polyene (step 11 followed by

    step 12), yielding 3CPC60. The transfer has been stud-

    ied both in toluene at ambient temperatures and in 2-

    methyltetrahydrofuran at lower temperatures. The energy

    transfer is an activated process (step 11), with

    Ea = 0.17 eV. This is consistent with transfer by a triplet

    energy-transfer relay, whereby energy first migrates from

    CP3C60 to the porphyrin (step 11), yielding C3PC60

    in a slow, thermally activated step. Rapid energy transfer

    from the porphyrin triplet to the carotenoid (step 12) gives

    the final state. Triplet relays of this general kind have been

    observed in photosynthetic reaction centres and are part

    of the system that protects the organism from damage bysinglet oxygen, whose production is sensitized by chloro-

    phyll triplet states (Gust et al. 1993a; Krasnovsky et al.

    1993).

    As explained above, the CPC60 state decays at

    77 K to give the carotenoid triplet. EPR investigation of

    the signal from this triplet state shows that the spin polar-

    ization of 3CPC60 is characteristic of a triplet formed by

    charge recombination of a singlet-derived radical pair.

    The kinetics of the decay of 3CPC60 to the ground state

    were also determined and the decay constants for the three

    triplet sublevels were measured. Similar spin states and

    dynamics, which are the signature of radical-pair recombi-

    nation reactions, have been observed in photosyntheticreaction centres but are unique to structure 10 (and close

    relatives) in the area of photosynthetic model systems

    (Carbonera et al. 1998). To aid our interpretation of the

    Phil. Trans. R. Soc. Lond. B (2002)

    EPR spectra of the triads, we carried out several other

    studies on the EPR properties of caroteneporphyrin

    dyads (Carbonera et al. 1997a,b). The generation in triad

    10 of a long-lived charge-separated state by photoinduced

    electron transfer, the low-temperature electron-transfer

    behaviour and the formation of a triplet state by charge

    recombination are phenomena heretofore observed mostly

    in photosynthetic reaction centres, rather than in modelsystems.

    Although triad 10 mimics some of the important fea-

    tures of photosynthetic electron transfer, the quantum

    yield of the final state is rather low (0.14). The initial

    CPC60 state is formed with a quantum yield

    approaching unity, but charge recombination to the

    ground state is more rapid than electron transfer from the

    carotenoid to yield the final state. In order to overcome

    this problem, triad 11 was designed. It differs from struc-

    ture 10 mainly in the nature of the carotenoid-to-porphyrin

    linkage (Kuciauskas et al. 2000). A major effect of this

    structural change is to increase the electron-donating

    ability of the carotenoid by 0.18 eV, although it alsoaffects the electronic coupling. In addition, the change

    raises the energy of CPC60 by 0.05 eV relative to

    structure 10. In accord with Marcus theory, these ener-

    getic changes increase the thermodynamic driving force

    and rate constant for formation of CPC60 from

    CPC60 (step 4). They also slightly increase the driv-

    ing force for charge recombination of CPC60 (step

    7). However, as this reaction occurs in the Marcus

    inverted region, the rate of recombination is slowed. The

    overall result is that the yield of CPC60 is 0.88 in

    structure 11 (see Appendix A).

    In order to retard the rate of charge recombination of

    CP

    C

    60 even more, and thereby increase the yield ofCPC60, triad 12 was synthesized (Bahr et al. 2000).

    In structure 12 (see Appendix A), the octaalkylporphyrin

    serving as the primary donor was replaced with a tetra-

  • 8/3/2019 Phil. Trans. R. Soc. Lond. 2002, 357, 1481

    8/18

    1488 T. A. Moore and others Artificial antenna systems and reaction centres

    phenyl-based porphyrin that increases the energy of P

    by ca. 200 mV. Therefore, in structure 12, the charge

    recombination process (step 7) from CPC60 is even

    further in the Marcus inverted region and thereby slower,

    allowing the forward reaction yielding the final charge-

    separated species CPC60 to better compete with it.

    Moreover, this change increases the driving force for hole

    transfer from the porphyrin radical cation to the carot-enoid (step 4), which is in the Marcus normal region and

    therefore is accelerated. The combination of a faster for-

    ward reaction combined with a slower recombination

    reaction from the intermediate CPC60 results in a

    measured quantum yield for the formation of CP

    C60 of ca. 100% in triad 12.

    The availability of structures 10, 11 and model com-

    pounds has allowed the investigation of the effects of tem-

    perature and solvent on the various electron-transfer steps.

    For example, electron-transfer dynamics on a very short

    time-scale as a function of solvent were studied in triad

    10 and in a related porphyrinfullerene dyad using sub-

    picosecond transient absorption techniques (Kuciauskas etal. 1998a). After porphyrin photoexcitation of the dyad,

    electron transfer to the fullerene was observed, yielding

    PC60. In toluene, this state recombines to give the ful-

    lerene first-excited singlet state, leading to fullerene singlet

    sensitization by an unusual two-step electron-transfer

    mechanism. This is somewhat analogous to the delayed

    fluorescence observed in natural reaction centres when

    secondary electron transfer is blocked. In polar solvents,

    structure 10 undergoes the electron-transfer process men-

    tioned previously to produce CPC60. The solvation

    dynamics of this state were investigated by monitoring car-

    otenoid radical cation absorption band bathochromic

    spectral shifts. Picosecond solvation processes were

    observed in six solvents, with solvation lifetimes ranging

    from 0.4 to 35 ps. In all solvents, electron-transfer rates

    were slower than solvation rates, but the difference is not

    large. It was also observed that when solvation of the caro-

    tenoid radical cation is faster, electron transfer is more

    rapid.

    An extensive comparison of the effects of temperature

    and solvent on triads 10 and 11 allowed several generaliza-

    tions (Kuciauskas et al. 2000). Rate constants for photoin-

    duced charge separation are much larger than those for

    charge recombination. In addition, the rate constants of

    the various electron-transfer reactions in structure 11

    depend only weakly on temperature and solvent. The yield

    of CPC60 in structure 11 is essentially independent

    of temperature down to 8 K. The rate constants for photo-

    induced electron transfer and for charge recombination of

    CPC60 do not change drastically when the solvent

    changes from a fluid solution to a rigid glass. Charge

    recombination of CPC60 also has some unusual fea-

    tures. In non-polar solvents or frozen glasses down to 8 K,

    charge recombination in structures 10 and 11 yields only

    the carotenoid triplet state. However, in polar solvents,

    charge recombination in structure 11 yields only the sing-

    let ground state (step 8). The rate of charge recombination

    of C

    PC

    60 in both structures 10 and 11 is temperature-dependent down to ca. 170 K, and temperature inde-

    pendent thereafter. The temperature-independent region

    is attributed to nuclear tunnelling.

    Phil. Trans. R. Soc. Lond. B (2002)

    (e) An artificial antennareaction centre complex:

    a molecular hexad

    Antenna function in photosynthesis has been mimicked

    in artificial systems of covalently linked arrays. For

    example, metalated and free base porphyrins have been

    linked with alkyne units to form dimers, trimers, tetra-

    mers, pentamers and more complex supermolecules that

    absorb light and rapidly and efficiently transfer singletexcitation to energy sinks such as a free base porphyrin

    moiety (Seth et al. 1994, 1996; Bothner-By et al. 1996;

    Hsiao et al. 1996; Wagner et al. 1996a; Strachan et al.

    1997; Lammi et al. 2000). Molecules of this type have

    been designed to act as photonic wires and gates

    (Wagner & Lindsey 1994; Wagner et al. 1996b). Other

    porphyrin arrays, some of them including more than one

    hundred porphyrin units, with demonstrated or potential

    antenna function have been reported (Van Patten et al.

    1998; Cho et al. 2000).

    Because of the progress made in the mimicry of both

    photosynthetic antennas and reaction centres, a logical

    next step would be to link an artificial light-harvestingarray with a synthetic reaction centre to produce a func-

    tional energy-conversion assembly. Indeed, a model pho-

    tosynthetic antenna consisting of four covalently linked

    zinc tetraarylporphyrins, (PZnP)3PZnC, covalently joined

    to a free base porphyrinfullerene artificial photosynthetic

    reaction centre, PC60, to form (PZnP)3PZnCPC60,hexad 13, has been reported (Kuciauskas et al. 1999).

    In order to evaluate the performance of hexad 13, rate

    constants were extracted from time-resolved spectroscopic

    data. The time-resolved fluorescence and absorption data

    yield a time constant of 700 ps that is associated with the

    zinc porphyrin moieties. The shortening of the zinc por-

    phyrin singlet lifetime from 2.4 ns in tetrad (PZnP)3PZnCto give a 700 ps component in a model pentad, (PZnP)3

    PZnCP and structure 13 (see Appendix A) is a result of

    singletsinglet energy transfer from the zinc porphyrins to

    the free base porphyrin. In structure 13, fluorescence from

    the free base porphyrin is quenched from its usual lifetime

    ofca. 10 ns to such a degree that fluorescence is no longer

    observable. The picosecond transient absorbance results

    yield additional time constants of 3 ps, 12 ps and 1.3 ns.

    The rate constant for photoinduced electron transfer

    and the charge recombination rate constant are directly

    observed experimentally. The reciprocal of the 3 ps time-

    constant detected in the transient absorption experiments

    is the rate constant for photoinduced electron transfer.This assignment is verified by the results for a model P

    C60 dyad, where the same value was obtained for the rate

    constant for photoinduced electron transfer. The charge

    recombination of (PZnP)3PZnCPC60 is associated

    with the 1.3 ns decay component observed in transient

    absorption, as demonstrated by the spectral stamp of the

    fullerene radical anion with absorption in the 1000 nm

    region. This lifetime is within a factor of 2.5 of the lifetime

    observed for the PC60 in a model dyad (480 ps).

    The values for the singlet energy-transfer rate constants

    between zinc porphyrins and between the central zinc por-

    phyrin and the free base porphyrin do not appear directly

    in the experimental decays. A kinetic model wasdeveloped to interpret the spectroscopic data and a sol-

    ution of a set of rate differential equations using the exper-

    imentally measured lifetimes of 12 ps and 700 ps as

  • 8/3/2019 Phil. Trans. R. Soc. Lond. 2002, 357, 1481

    9/18

    Artificial antenna systems and reaction centres T. A. Moore and others 1489

    constants yields values of 50 ps for the rate constant for

    energy transfer between metal porphyrins and 240 ps for

    the rate constant for energy transfer between the zinc and

    free base porphyrins. The value thus obtained for the rate

    constant for energy transfer between metal porphyrins is

    in excellent agreement with the results of Hsiao and

    coworkers who measured a rate constant of 53 ps for sing-

    let energy transfer between two zinc porphyrins in a lineartrimeric array of zinc porphyrins with diphenylethyne link-

    ages very similar to those in structure 13 (Hsiao et al.

    1996).

    The singletsinglet energy transfer between the central

    zinc porphyrin in the antenna of structure 13 and the free

    base porphyrin is slow (240 ps) compared with that

    between the zinc porphyrins (50 ps). Thus, the energy

    transfer between the central zinc porphyrin and the free

    base porphyrin creates a bottleneck that limits the quan-

    tum yield of the final charge-separated state (PZnP)3PZnC

    PC60.

    The quantum yield of (PZnP)3PZnCPC60 is wave-

    length dependent. Based on light absorbed by the freebase porphyrin, the yield is unity, due to the very large

    rate constant for photoinduced electron transfer. The

    quantum yield based on excitation of the zinc porphyrins

    in the antennas is 0.70. In this process, the light-gathering

    power of the system is increased considerably at many

    wavelengths, as four zinc porphyrin moieties feed exci-

    tation energy to the reaction centre. In hexad 14 structural

    modifications to increase the electronic coupling between

    PZnCP and the thermodynamic driving force for selected

    electron-transfer steps were carried out to improve the

    performance of this device in terms of lifetime and quan-

    tum yield of the final charge-separated state (Kodis et al.

    2002).Hexad 14 differs from structure 13 in several important

    ways. By changing the primary donor porphyrin from the

    free base octaalkylporphyrin of structure 13 to a free base

    tetraphenyl-type porphyrin, the energy of the correspond-

    ing radical cation is increased by ca. 200 mV. This pro-

    vides an additional thermodynamic driving force for hole

    transfer from the primary donor to the Zn porphyrins act-

    ing as the antenna. Evidence for this is provided by the

    appearance in the transient spectra of the Zn porphyrin

    radical cation in 380 ps and greatly increased lifetime of

    the charge-separated state in structure 14 (see Appendix

    A) of up to 24 s (solvent dependent) versus 1.3 ns in

    structure 13. As mentioned above, this change of severalorders of magnitude in lifetime is characteristic of multi-

    component systems in which a secondary electron-transfer

    step results in formation of a neutral chromophore separ-

    ating electron-hole pairs. Increasing the energy of the por-

    phyrin radical cation also lowers the driving force for

    photoinduced electron transfer from the primary donor to

    the C60 and, indeed, charge separation is slower, 25 ps in

    structure 14 versus 3 ps in structure 13, because this step

    is in the Marcus normal region. Twenty-five picoseconds

    is still sufficiently fast so that the yield of the photoinduced

    electron-transfer process, competing against fluorescence

    and intersystem crossing, is not measurably changed

    from 100%.A second consequence of the change to a tetraphenyl-

    based porphyrin as the primary donor in hexad 14 is that

    there are no methyl groups at the beta positions. This

    Phil. Trans. R. Soc. Lond. B (2002)

    allows a more coplanar conformation between the central

    Zn porphyrin of the antenna and the primary donor,

    which increases the electronic coupling and results in

    much faster singlet energy transfer. The time for this step,

    which was 240 ps in structure 13 and limited the overall

    quantum yield of energy transfer, is reduced to 30 ps in

    structure 14, giving an energy-transfer yield of unity.

    Therefore, the quantum yield of photoinduced electrontransfer from the primary donor to the C60 in structure

    14 is essentially 100%, regardless of which porphyrin is

    initially excited.

    In addition to the changes noted above in thermodyn-

    amics and conformation that are brought about by chang-

    ing the primary donor from an octaalkylporphyrin to a

    tetraphenyl-based porphyrin, there are electronic effects

    arising from changes in the molecular orbital nodal pattern

    of the two porphyrins. Specifically relevant to hole transfer

    from the primary donor to the Zn porphyrin of the

    antenna, the HOMO of the octaalkylporphyrin has a node

    at the meso positions whereas the HOMO of the

    tetraphenyl-based porphyrin has amplitude at the mesopositions. Because the meso position is the attachment

    point of the Zn porphyrin array to the primary donor por-

    phyrin, and because the HOMOs are involved in hole

    transfer, this aspect of the electronic-coupling term should

    be larger in structure 14 than in structure 13. The result of

    improved thermodynamics for hole transfer and increased

    electronic coupling due to conformation and orbital

    effects is hole transfer from the radical cation of the pri-

    mary donor to the antenna in 380 ps. This is sufficiently

    fast for this step to have a yield of close to unity so that

    the overall quantum yield of long-lived charge-separated

    species in hexad 14 is near 100%.

    6. ARTIFICIAL PHOTOSYNTHETIC MEMBRANES

    Molecular triad 15, which consists of a porphyrin chro-

    mophore (P) bound covalently to a naphthoquinone

    derivative (NQ) and a carotenoid electron donor (C), was

    designed both to undergo multistep electron transfer and

    to organize itself in a bilayer lipid membrane. These two

    characteristics are discussed below.

    Excitation of the porphyrin moiety of structure 15 (see

    Appendix A) (the photophysics of structure 15 is not

    strongly dependent on the ionization state of the car-

    boxylate group) leads to C1PNQ, which decays by pho-

    toinduced electron transfer to give CP

    NQ

    with aquantum yield near unity. Competing with charge recom-

    bination, a second electron transfer from the carotenoid

    to the porphyrin radical cation occurs, yielding a C P

    NQ charge-separated state. In this state, the radicals are

    separated by the neutral porphyrin moiety and therefore

    charge recombination is slower by several orders of magnitude

    than that of a comparable dyad. We have used this strategy

    in a variety of other reaction-centre mimics (vide supra),

    some of which rival the natural systems in terms of quan-

    tum yield, fraction of energy stored and lifetime for charge

    separation (Gust & Moore 1991; Gust et al. 1993b).

    It has been demonstrated that, under certain conditions,

    triad 15 inserts unidirectionally into a liposomal phospho-lipid bilayer membrane (Steinberg-Yfrach et al. 1997).

    How can this observation be rationalized? Triad 15 is

    highly amphiphilic due to the hydrophobic carotenoid pig-

  • 8/3/2019 Phil. Trans. R. Soc. Lond. 2002, 357, 1481

    10/18

    1490 T. A. Moore and others Artificial antenna systems and reaction centres

    C

    C C

    C

    C

    C

    C

    C

    6

    1

    Q s H

    7

    QsH+

    Qs

    5

    Qs

    Qs

    +

    Qsh

    v

    +

    +

    _ 2

    +

    Q s

    3

    +

    Q s H

    +

    4

    P P

    H

    QQ

    H2O, buffer

    Q

    P

    P

    P

    P

    Q

    Q

    Q

    Q

    P

    P

    Q

    _

    H

    pyraninetrisulphonate

    9-aminoacridine

    8-anilinonaphthalene-1-sulphonic

    acid

    H2O,

    or

    or

    buffer

    .

    .

    ...

    .

    .

    .

    Figure 4. Steps 17 illustrate a minimum set of proposed

    elementary processes involved in the redox loop-based

    proton translocation across a liposomal bilayer membrane.

    ment distal to the carboxylate group that is attached to

    the naphthoquinone. When a small amount of structure

    15 (dissolved in a suitable solvent such as tetrahydrofuran)

    is added to an aqueous solution of liposomes, the amphi-

    philic molecules associate with the membrane. It is prob-

    able that the most stable association features thehydrophobic carotenoid moiety in the low-dielectric

    interior of the bilayer and the hydrophilic carboxylate

    group at or near the aqueous interface. Because structure

    15 is added from the bulk aqueous phase on the outside

    of the liposomal bilayer, the expected initial arrangement

    would be vectorial, with the majority of triads having their

    carboxylate-bearing naphthoquinones near the bulk aque-

    ous phase as shown schematically in figure 4. The ener-

    getic cost of transporting the carboxylate across the bilayer

    would slow, but not prevent, flip-flop diffusion that would

    ultimately randomize the orientation.

    (a) Artificial proton pumpThe availability of artificial reaction centres that self-

    assemble vectorially into bilayer lipid membranes makes

    it possible to design a proton pump that will generate a

    transmembrane gradient in proton electrochemical poten-

    tial, the next step in converting solar energy to a biologi-

    cally useful form (Steinberg-Yfrach et al. 1997). In order

    to couple the vectorial redox potential in the species C

    PNQ to proton translocation, a lipophilic shuttle quin-

    one, Qs, was incorporated into the liposomal bilayer where

    it functions as a redox loop to shuttle protons across the

    bilayer. A proposed mechanism for this proton translo-

    cation is shown diagrammatically in figure 4.

    To begin with, excitation of structure 15 generates theCPNQ charge-separated state via the multistep

    electron-transfer process described above. The shuttle

    quinone Qs is more easily reduced than the naphthoqui-

    Phil. Trans. R. Soc. Lond. B (2002)

    none of the triad and therefore accepts an electron from

    the charge-separated state to generate the Qs radical

    anion, which is basic enough to accept a proton from the

    nearby external aqueous phase. The resulting neutral

    semiquinone is lipid soluble and diffuses within the lipo-

    philic part of the bilayer. When it encounters a carotenoid

    radical cation near the internal aqueous phase, it is readily

    oxidized, yielding a protonated quinone, QsH

    . Thisstrong acid releases a proton to the nearby interior aque-

    ous volume, thereby lowering the pH inside the liposome

    and regenerating the components of the proton pump. In

    other words, the vectorial redox potential of CPNQ

    drives a redox loop in which the reduced (protonated)

    form of Qs carries protons inwards and the oxidized

    (unprotonated) form of Qs completes the loop by equilib-

    rating across the membrane. The specific details of the

    proton-pumping photocycle may differ somewhat from

    this simple picture, but figure 4 describes the basic fea-

    tures of the system.

    Proton import has been monitored by several methods.

    A pH-sensitive fluorescent dye, pyraninetrisulphonate, wastrapped inside the liposomes, or 8-aminoacridine was used

    to measure pH directly, or 8-anilinonaphthalene-1-

    sulfonic acid was used to measure the transmembrane

    electrical potential . Illumination of the liposomes with

    light absorbed by the porphyrin moiety of structure 15

    leads to a change in the fluorescence properties of these

    dyes that signals import of protons into the liposomes. A

    H of ca. 4.4 kcal mole1 comprised of pH ca. 2.1

    and ca. 70 mV is attained after a few minutes of

    irradiation with less than 0.1 mW of red light. The pH and

    gradients thus generated can be relaxed by addition of

    an ionophore such as FCCP. The photochemical import

    of protons is enhanced by the addition of potassium andvalinomycin, which presumably reduce the uncompen-

    sated charge buildup () in the intraliposomal volume.

    Conversely, increasing the buffer concentration inside the

    liposomes decreases the magnitude ofpH and results in

    the generation of a greater . Thus, a photocyclic trans-

    membrane proton pump has been assembled.

    (b) ATP synthesis by the artificial membrane

    Proton motive force is a basic form of biological energy

    that is used to drive most energy-requiring reactions in

    living organisms. Most importantly, it is used to power the

    synthesis of ATP. Having prepared a light-driven trans-

    membrane proton-pumping system, we have used it topower ATP synthesis in a biomimetic fashion. In nature,

    ATP is synthesized by F0F1ATP synthase, an enzymatic

    system that spans the lipid bilayer. Protons pass through

    the enzyme from one side of the membrane to the other

    in the direction of the pH gradient and the energy pro-

    vided by relaxation of the electrochemical potential gradi-

    ent is used to synthesize ATP from ADP and Pi.

    Conditions for extraction of the CF0F1ATP synthase

    from spinach chloroplast thylakoids and reconstitution of

    the functional enzyme into liposomes have been

    developed. Using this methodology, the enzyme has been

    reconstituted into liposomes containing the components

    of the light-driven proton pump discussed above(Steinberg-Yfrach et al. 1998).

    The system is shown as a diagram in figure 5. The lipo-

    somes were formed from a 10 : 1 mixture of egg phosphat-

  • 8/3/2019 Phil. Trans. R. Soc. Lond. 2002, 357, 1481

    11/18

    Artificial antenna systems and reaction centres T. A. Moore and others 1491

    OO

    O

    OO

    O

    .

    triad 15

    H+

    Q

    P

    C

    P

    Q

    C

    hv

    QS

    +

    .HQs

    ._

    H+

    4H+

    ADP + Pi

    Fo

    F1

    4H+

    ATP

    CF0F1-ATP synthase

    intraliposomal volume

    Figure 5. Schematic diagram of CFoF1 ATP synthase inserted into a liposomal membrane containing the components of the

    artificial reaction centre and redox loop-based proton pump. The enzyme is reconstituted into the membrane with the

    nucleotide binding sites in the external aqueous phase.

    idylcholine and egg phosphatidic acid containing 20%

    cholesterol. The ATP synthase was inserted into the

    bilayer with the ATP-synthesizing portion of the enzyme

    extending out into the external aqueous solution. The

    constituents of the proton pump were incorporated as

    described above. Thioredoxin (10 M) was added to acti-

    vate the enzyme and ADP, Pi and an initial amount of

    ATP were added to the external phase. The aqueous

    phases were adjusted to a pH of 8 prior to illumination.

    The liposomes were illuminated for various periods of

    time with laser light at 633 nm, which was absorbed only

    by the triad. Illumination was terminated, and the ATP

    produced was measured using a calibrated luciferin/luciferase assay, which is based on the luminescence of

    oxyluciferin. The amount of ATP produced during each

    period of illumination was determined; typical results are

    presented graphically in figure 6.

    In order to establish that this increase was indeed due

    to ATP synthesis driven by light-induced proton motive

    force, several control experiments were performed. As

    shown in figure 6, no ATP was synthesized when 1 M

    FCCP, the proton ionophore mentioned above, was

    added to the system, making the membrane permeable to

    protons. Addition of 2 M tentoxin, which is a specific

    inhibitor of CF0F1ATP synthase, halted ATP synthesis,

    as did omission of the shuttle quinone Qs or ADP. Like-wise, omission of Pi, triad 15 or ATP synthase abolished

    the effect.

    At low levels of illumination, where ATP synthesis was

    limited by light absorption, it was found that one ATP

    molecule was synthesized for every 14 photons incident

    on the sample. As ca. 50% of the light was absorbed by

    the scattering liposome solutions, this corresponds to a

    quantum yield of ca. 0.15. Assuming that four protons

    must be transported out of the liposome by ATP synthase

    per molecule of ATP produced, the yield of the proton

    pump must be ca. 0.6. It should be noted that the lipo-

    somes used for ATP synthesis were of a different size and

    lipid mix from those used in the proton translocationexperiments discussed in the previous section ( 6a), and

    that ca. 1000 triad 15 molecules could be incorporated

    into one proteoliposome. In the proton translocation

    Phil. Trans. R. Soc. Lond. B (2002)

    ATP

    produced(molATP1

    011)

    Figure 6. The synthesis of ATP as a function of illumination

    time at different ratios of [ATP]/[ADP] (triangles, [ATP] =

    [ADP] = 0.2 mM and [Pi] = 5 mM; filled circles, [ATP] =

    0.2 mM, [ADP] = 0.02 mM and [Pi] = 5 mM. Also shown

    are results for control experiments showing that ATP was

    not synthesized (open circles, 1 M FCCP; filled triangles,2 M tentoxin; diamonds, [Qs] = 0; open squares, [ADP] = 0.

    experiments, only ca. 40 triad molecules were present in

    each liposome.

    As illustrated in figure 6, the liposome system can syn-

    thesize ATP against a considerable ATP chemical poten-

    tial. Under the conditions of 0.2 mM ATP, 0.02 mM

    ADP and 5 mM Pi shown in figure 6, ATP is synthesized

    against a chemical potential of ca. 12 kcal mol1. Thus,

    the system is converting light energy to chemical potential,

    rather than simply catalyzing an exergonic reaction. Based

    on the ATP potential, the quantum yield and the energy

    of 633 nm photons, we estimate that ca. 4% of theabsorbed light energy is conserved in the system. The lim-

    its of the system have not as yet been reached, nor has the

    system been optimized.

  • 8/3/2019 Phil. Trans. R. Soc. Lond. 2002, 357, 1481

    12/18

    1492 T. A. Moore and others Artificial antenna systems and reaction centres

    7. CONCLUSION

    Supramolecular structures that mimic the antenna,

    reaction centre and proton pumps of natural photosyn-

    thetic membranes can be designed and synthesized. Arti-

    ficial reaction centres demonstrate that rapid charge

    separation and relatively slow charge recombination

    results in relatively long-lived charge-separated states that

    facilitate the use of the energy stored in these states in

    homogeneous solution, or even in interfacial reactions.

    The ability to carry out photoinduced electron transfer in

    media of low dielectric constant, in rigid media and/or at

    low temperatures indicates that fullerene-based systems

    would be good candidates for applications such as mol-

    ecular scale (opto)electronics, sensors or bioenergetics

    where photochemistry would have to occur in self-

    assembled or LangmuirBlodgett monolayers, in plastics

    or gels, in biological or artificial membranes, in micelles,

    etc. Large arrays of porphyrin-based antennas can be con-

    structed in order to harvest light efficiently. If these arrays

    are coupled to reaction centres based on fullerenes such

    N

    NN

    N

    H

    H

    C

    O

    N

    H

    Structure 1. (Carotenopurpurin.)

    N

    NN

    N

    H

    H

    C

    O

    OCH 3

    Structure 2. (Purpurin.)

    Phil. Trans. R. Soc. Lond. B (2002)

    as in hexad 14, supermolecules can be assembled that can

    be effective transducers of light energy to potential energy

    in the form of charge-separated states. Furthermore, it is

    possible to mimic the basic features of energy conversion

    in bacterial photosynthetic membranes in a mostly arti-

    ficial, bionic construct that employs artificial reaction

    centres. Optimization of such systems could provide solar

    biological power packs that might be used to drive theenzymatic synthesis of other high-energy or high-value

    compounds, or power natural or artificial nanoscale

    machines. Both scientific and practical uses of such mod-

    ules can be envisioned.

    This work was supported by grants from the Department of

    Energy (DE-FG03-93ER14404) and the National Science

    Foundation (CHE-0078835). This is publication 538 from the

    ASU Centre for the Study of Early Events in Photosynthesis.

    APPENDIX A

    Structures 115.

  • 8/3/2019 Phil. Trans. R. Soc. Lond. 2002, 357, 1481

    13/18

    Artificial antenna systems and reaction centres T. A. Moore and others 1493

    N

    H

    C

    O

    H3C

    Structure 3. (Carotenoid of structure 1.)

    N

    N

    N

    N

    N

    N

    NN Zn

    O

    O O

    O

    N

    O

    O

    H

    C

    O

    Structure 4. (Carotenophthalocyanine.)

    N

    N

    N

    N

    N

    N

    NN Zn

    O

    O O

    O

    N

    O

    O

    H

    C

    O

    Structure 5. (Phthalocyanine.)

    C

    O

    N

    H

    Structure 6. (Carotenoid of structure 2.)

    Phil. Trans. R. Soc. Lond. B (2002)

    C

    O

    N

    H

    H

    Structure 7.

    NH

    NH

    N

    N

    Structure 8.

  • 8/3/2019 Phil. Trans. R. Soc. Lond. 2002, 357, 1481

    14/18

    1494 T. A. Moore and others Artificial antenna systems and reaction centres

    NH3C

    NH

    NH

    N

    N

    Structure 9.

    CN

    N N

    N N

    HH

    O

    N

    H

    Structure 10.

    NNH

    HH

    C

    O

    N N

    N N

    Structure 11.

    NN

    N N

    N N

    HH

    H

    C

    O

    Structure 12.

    Phil. Trans. R. Soc. Lond. B (2002)

  • 8/3/2019 Phil. Trans. R. Soc. Lond. 2002, 357, 1481

    15/18

    Artificial antenna systems and reaction centres T. A. Moore and others 1495

    NH

    HN

    N

    N

    N

    N

    N

    N

    N

    N

    N

    N

    N

    N

    N

    N

    N

    N

    N

    N

    N

    Zn

    Zn

    Zn

    Zn

    Structure 13.

    N

    N

    N

    N

    HN

    N

    N

    N

    N

    N

    N

    N

    N

    N

    N

    N

    N

    N

    N

    N

    N

    Zn

    Zn

    Zn

    Zn H

    Structure 14.

    Phil. Trans. R. Soc. Lond. B (2002)

  • 8/3/2019 Phil. Trans. R. Soc. Lond. 2002, 357, 1481

    16/18

    1496 T. A. Moore and others Artificial antenna systems and reaction centres

    N

    N

    N

    NN

    O

    O

    CN

    H H

    O

    C

    O

    HH C

    O

    O-

    Structure 15.

    REFERENCES

    Bahr, J. L., Kuciauskas, D., Liddell, P. A., Moore, A. L.,

    Moore, T. A. & Gust, D. 2000 Driving force and electronic

    coupling effects on photoinduced electron transfer in a

    fullerene-based molecular triad. Photochem. Photobiol. 72,598611.

    Bensasson, R. L., Land, E. J. & Maudinas, B. 1976 Triplet

    states of carotenoids from photosynthetic bacteria studied by

    nanosecond ultraviolet and electron pulse irradiation.

    Photochem. Photobiol. 23, 189193.

    Bothner-By, A. A., Dadok, J., Johnson, T. E. & Lindsey, J. S.

    1996 Molecular dynamics of covalently-linked multi-por-

    phyrin arrays. J. Phys. Chem. 100, 17 55117 557.

    Carbonera, D., Di Valentin, M., Corvaja, C., Giacometti, G.,

    Agostini, G., Liddell, P. A., Moore, A. L., Moore, T. A. &

    Gust, D. 1997a Carotenoid triplet detection by time

    resolved EPR spectroscopy in carotenopyropheophorbide

    dyads. J. Photochem. Photobiol. A. Chem. 105, 329335.

    Carbonera, D., De Valentin, M., Agostini, G., Giacometti, G.,

    Liddell, P. A., Gust, D., Moore, A. L. & Moore, T. A.

    1997b Energy transfer and spin polarization of the carot-

    enoid triplet state in synthetic carotenoporphyrin dyads and

    in natural antenna complexes. Appl. Magn. Reson. 13,

    487504.

    Carbonera, D., Di Valentin, M., Corvaja, C., Agostini, G.,

    Giacometti, G., Liddell, P. A., Kuciauskas, D., Moore,

    A. L., Moore, T. A. & Gust, D. 1998 EPR investigation of

    photoinduced radical pair formation and decay to a triplet

    state in a caroteneporphyrinfullerene triad. J. Am. Chem.

    Soc. 120, 43984405.

    Cardoso, S. L., Nicodem, D. E., Moore, T. A., Moore,

    A. L. & Gust, D. 1996 Synthesis and fluorescence quench-

    ing studies of a series of carotenoporphyrins with carotenoids

    of various lengths. J. Braz. Chem. Soc. 7, 1929.Cho, H. S., Song, N. W., Kim, Y. H., Jeoung, S.-Ch., Hahn,

    S., Kim, D., Kim, S. K., Yoshida, N. & Osuka, A. 2000

    Ultrafast energy relaxation dynamics of directly linked por-

    phyrin arrays. J. Phys. Chem. A 104, 32873298.

    Cogdell, R. J. & Frank, H. A. 1993 Photochemistry and func-

    tion of carotenoids in photosynthesis. In Carotenoids in photo-

    synthesis (ed. A. Young & G. Britton), pp. 252326. London:

    Chapman & Hall.

    Cohen-Bazire, G. & Stanier, R. Y. 1958 Inhibition of carot-

    enoid synthesis in photosynthetic bacteria. Nature 181,

    250252.

    Demmig-Adams, B. 1992 Operation of the xanthophyll cycle

    in higher plants in response to diurnal changes in incident

    sunlight. Planta 186, 390398.Foote, C. S., Chang, Y. C. & Denny, R. W. 1970 Chemistry of

    singlet oxygen. X. Carotenoid quenching parallels biological

    protection. J. Am. Chem. Soc. 92, 52165218.

    Phil. Trans. R. Soc. Lond. B (2002)

    Forster, T. 1948 Intermolecular energy transfer and fluor-

    escence. Ann. Phys. 2, 5575.

    Frank, H. A. & Cogdell, R. J. 1996 Carotenoid in photosyn-

    thesis. Photochem. Photobiol. 63, 257264.

    Frank, H., Cua, A., Chynwat, V., Young, A., Gosztola, D. &

    Wasielewski, M. R. 1996 The lifetimes and energies of thefirst excited singlet states of diadinoxanthin and diatoxan-

    thin: the role of these molecules in excess energy dissipation

    in algae. Biochim. Biophys. Acta 1277, 243252.

    Guldi, D. M. 2000 Fullerenes: three dimensional electron

    acceptor materials. J. Chem. Soc. Chem. Commun. 321327.

    Gust, D. & Moore, T. A. 1991 Mimicking photosynthetic elec-

    tron and energy transfer. Adv. Photochemistry 16, 165.

    Gust, D. & Moore, T. A. 2000 Intramolecular photoinduced

    electron transfer reactions of porphyrins. In The porphyrin

    handbook, vol 8. Electron transfer (ed. K. M. Kadish, K. M.

    Smith & R. Guillard), pp. 153190. San Diego, CA: Aca-

    demic.

    Gust, D., Moore, T. A., Moore, A. L., Devadoss, C., Liddell,

    P. A., Hermant, R., Nieman, R. A., Demanche, L. J.,DeGraziano, J. M. & Gouni, I. 1992 Triplet and singlet

    energy transfer in caroteneporphyrin dyads: the role of the

    linkage bonds. J. Am. Chem. Soc. 114, 35903603.

    Gust, D., Moore, T. A., Moore, A. L., Krasnovsky Jr, A. A.,

    Liddell, P. A., Nicodem, D., DeGraziano, J. M., Kerrigan,

    P., Makings, L. R. & Pessiki, P. J. 1993a Mimicking the

    photosynthetic triplet energy transfer relay. J. Am. Chem.

    Soc. 115, 56845691.

    Gust, D., Moore, T. A. & Moore, A. L. 1993b Molecular mim-

    icry of photosynthetic energy and electron transfer. Accounts

    Chem. Res. 26, 198205.

    Gust, D., Moore, T. A., Moore, A. L., Liddell, P. A., Kuc-

    iauskas, D., Sumida, J. P., Nash, B. & Nguyen, D. 1997 A

    caroteneporphyrinfullerene triad: photoinduced chargeseparation and charge recombination to a triplet state. In

    Recent advances in the chemistry and physics of fullerenes and

    related materials, vol. 4 (ed. K. M. Kadish & R. S. Ruoff),

    pp. 924. Pennington, NJ: The Electrochemical Society.

    Gust, D., Moore, T. A., Moore, A. L., Kuciauskas, D.,

    Liddell, P. A. & Halbert, B. D. 1998 Mimicry of carotenoid

    photoprotection in artificial photosynthetic reaction centers:

    triplettriplet energy transfer by a relay mechanism. J. Photo-

    chem. Photobiol. B 43, 209216.

    Gust, D., Moore, T. A. & Moore, A. L. 1999 An arti ficial pho-

    tosynthetic membrane. Zeitschrift fur Physikalische Chemie Bd.

    213, S. 149155.

    Gust, D., Moore, T. A. & Moore, A. L. 2000 Photochemistry

    of supramolecular systems containing C60. J. Photochem.Photobiol. B 58, 6371.

    Gust, D., Moore, A. L. & Moore, T. A. 2001a Covalently

    linked systems containing porphyrin units. In Electron transfer

  • 8/3/2019 Phil. Trans. R. Soc. Lond. 2002, 357, 1481

    17/18

    Artificial antenna systems and reaction centres T. A. Moore and others 1497

    in chemistry, vol. 3. Biological and artificial supramolecular sys-

    tems (ed. V. Balzani), pp. 272336. Weinheim, Germany:

    Wiley-VCH.

    Gust, D., Moore, T. A. & Moore, A. L. 2001b Mimicking pho-

    tosynthetic solar energy transduction. Accounts Chem. Res.

    34, 4048.

    Hasharoni, K., Levanon, H., Tang, J., Bowman, M. K.,

    Norris, J. R., Gust, D., Moore, T. A. & Moore, A. L. 1990

    Singlet photochemistry in model photosynthesis: identifi-cation of charge separated intermediates by Fourier trans-

    form and CW EPR spectroscopies. J. Am. Chem. Soc. 112,

    64776481.

    Hermant, R. M., Liddell, P. A., Lin, S., Alden, R. G., Kang,

    H. K., Moore, A. L., Moore, T. A. & Gust, D. 1993 Mim-

    icking carotenoid quenching of chlorophyll fluorescence. J.

    Am. Chem. Soc. 115, 20802081.

    Horton, P., Ruban, A. V. & Walters, R. G. 1996 Regulation

    of light harvesting in green plants. A. Rev. Plant Physiol.

    Plant Mol. Biol. 47, 655684.

    Hsiao, J.-S., Krueger, B. P., Wagner, R. W., Johnson, T. E.,

    Delaney, J. K., Mauzerall, D. C., Fleming, G. R., Lindsey,

    J. S., Bocian, D. F. & Donohoe, R. J. 1996 Soluble multi-

    porphyrin arrays. 2. Photodynamics of energy-transfer pro-cesses. J. Am. Chem. Soc. 118, 11 18111 193.

    Imahori, H. (and 11 others) 1995 Photoinduced electron

    transfer in a carotenobuckminsterfullerene dyad. Photochem.

    Photobiol. 62, 10091014.

    Imahori, H., Hagiwara, K., Akiyama, T., Aoki, M., Taniguchi,

    S., Okada, T., Shirakawa, M. & Sakata, Y. 1996 The small

    reorganization energy of C60 in electron transfer. Chem.

    Phys. Lett. 263, 545550.

    Kandori, H., Sansabe, H. & Mimuro, M. 1994 Direct determi-

    nation of the lifetime of the S2 state of beta-carotene by femto-

    second time-resolved fluorescence spectroscopy. J. Am.

    Chem. Soc. 116, 26712672.

    Kodis, G., Liddell, P. A., de la Garza, L., Clausen, P. C.,

    Lindsey, J. S., Moore, A. L., Moore, T. A. & Gust, D. 2002Efficient energy transfer and electron transfer in an artificial

    photosynthetic antennareaction center complex. J. Phys.

    Chem. A 106, 20362048.

    Kohler, B. E. 1991 Electronic properties of linear polyenes. In

    Conjugated polymers (ed. J. L. Bredas & R. Silbey), pp. 405

    434. Dordrecht, The Netherlands: Kluwer.

    Koyama, Y., Kuki, M., Andersson, P. O. & Gillbro, T. 1996

    Singlet excited states and the light-harvesting function of

    carotenoids in bacterial photosynthesis. Photochem. Photobiol.

    63, 243256.

    Krasnovsky Jr, A. A., Cheng, P., Blankenship, R. E., Moore,

    T. A. & Gust, D. 1993 The photophysics of monomeric bac-

    teriochlorophylls c and d and their derivatives: properties of

    the triplet state and singlet oxygen photogeneration andquenching. Photochem. Photobiol. 57, 324330.

    Krinsky, N. I. 1979 Carotenoid protection against oxidation.

    Pure Appl. Chem. 51, 649660.

    Krueger, B. P., Scholes, G. D., Jimenez, R. & Fleming, G. R.

    1998 Electronic excitation transfer from carotenoid to bac-

    teriochlorophyl in the purple bacterium Rhodopseudomonas

    acidophila. J. Phys. Chem. 102, 22842292.

    Kuciauskas, D., Lin, S., Seely, G. R., Moore, A. L., Moore,

    T. A., Gust, D., Drovetskaya, T., Reed, C. A. & Boyd,

    P. D. W. 1996 Energy and photoinduced electron transfer

    in porphyrin-fullerene dyads. J. Phys. Chem. 100, 15 926

    15 932.

    Kuciauskas, D., Liddell, P. A., Moore, T. A., Moore, A. L. &

    Gust, D. 1998a Solvent effects and electron transfer dynam-ics in a porphyrin-fullerene dyad and a carotenoporphyrin-

    fullerene triad. In Recent advances in the chemistry and physics

    of fullerenes and related materials, vol 6 (ed. K. M. Kadish &

    Phil. Trans. R. Soc. Lond. B (2002)

    R. S. Ruoff), pp. 242261. Pennington, NJ: The Electro-

    chemical Society.

    Kuciauskas, D., Liddell, P. A., Moore, A. L., Moore, T. A. &

    Gust, D. 1998b Magnetic switching of charge separation life-

    times in artificial photosynthetic reaction centers. J. Am.

    Chem. Soc. 120, 10 88010 886.

    Kuciauskas, D., Liddell, P. A., Lin, S., Johnson, T. E.,

    Weghorn, S. J., Lindsey, J. S., Moore, A. L., Moore, T. A. &

    Gust, D. 1999 An artificial photosynthetic antennareactioncenter complex. J. Am. Chem. Soc. 121, 86048614.

    Kuciauskas, D., Liddell, P. A., Lin, S., Stone, S. G., Moore,

    A. L., Moore, T. A. & Gust, D. 2000 Photoinduced electron

    transfer in carotenoporphyrinfullerene triads: temperature

    and solvent effects. J. Phys. Chem. 140, 43074321.

    Lammi, R. K., Ambroise, A., Balasubramanian, T., Wagner,

    R. W., Bocian, D. F., Holten, D. & Lindsey, J. S. 2000

    Structural control of photoinduced energy transfer between

    adjacent and distant sites in multiporphyrin arrays. J. Am.

    Chem. Soc. 122, 75797591.

    Lewis, J. E., Moore, T. A., Benin, D., Gust, D., Nicodem,

    D. & Nonell, S. 1994 The triplet energy of a carotenoid pig-

    ment determined by photoacoustic calorimetry. Photochem.

    Photobiol. 59S, 35S.Liddell, P. A., Sumida, J. P., Macpherson, A. N., Noss, L.,

    Seely, G. R., Clark, K. N., Moore, A. L., Moore, T. A. &

    Gust, D. 1994 Preparation and photophysical studies of

    porphyrin-C60 dyads. Photochem. Photobiol. 60, 537541.

    Liddell, P. A., Kuciauskas, D., Sumida, J. P., Nash, B.,

    Nguyen, D., Moore, A. L., Moore, T. A. & Gust, D. 1997

    Photoinduced charge separation and charge recombination

    to a triplet state in a caroteneporphyrinfullerene triad. J.

    Am. Chem. Soc. 119, 14001405.

    Luo, C., Guldi, D. M., Imahori, H., Tamaki, K. & Sakata, Y.

    2000 Sequential energy and electron transfer in an artificial

    reaction center: formation of a long-lived charge-separated

    state. J. Am. Chem. Soc. 122, 65356551.

    Macpherson, A. N. & Gillbro, T. 1998 Solvent dependence of

    the ultrafast S2S1 internal conversion rate of-carotene. J.

    Phys. Chem. 102, 50495058.

    Macpherson, A. N., Liddell, P. A., Kuciauskas, D., Tatman,

    D., Gillbro, T., Gust, D., Moore, T. A. & Moore, A. L.

    2002 Ultrafast energy transfer from a carotenoid to a chlorin

    in a simple artificial photosynthetic antenna. J. Phys.

    Chem. (Submitted.)

    Mathis, P. & Kleo, J. 1973 The triplet state of -carotene and

    of analog polyenes of different length. Photochem. Photobiol.

    18, 343346.

    Moore, T. A. (and 11 others) 1984 Photodriven charge separ-

    ation in a carotenoporphyrin-quinone triad. Nature 307,

    630632.

    Moore, T. A., Gust, D. & Moore, A. L. 1990 The function

    of carotenoid pigments in photosynthesis and their possibleinvolvement in the evolution of higher plants. In Carotenoids:

    chemistry and biology (ed. N. I. Krinsky, M. M. Mathews-

    Roth & R. F. Taylor), pp. 223228. New York: Plenum.

    Moore, T. A., Gust, D. & Moore, A. L. 1994 Carotenoids:

    natures unique pigments for light and energy processing.

    Pure Appl. Chem. 66, 10331040.

    Quinones, M. A., Lu, Z. & Zeiger, E. 1996 Close correspon-

    dence between the action spectra for the blue light responses

    of the guard cell and coleoptile chloroplasts, and the spectra

    for blue light-dependent stomatal opening and coleoptile

    phototropism. Proc. Natl Acad. Sci. USA 93, 22242228.

    Ricci, M., Bradforth, S. E., Jimenez, R. & Fleming, G. R. 1996

    Internal conversion and energy transfer dynamics of sphero-

    idene in solution and in the LH-1 and LH-2 light-harvestingcomplexes. Chem. Phys. Lett. 259, 381390.

    Seth, J., Palaniappan, V., Johnson, T. E., Prathapan, S., Lind-

    sey, J. S. & Bocian, D. F. 1994 Investigation of electronic

  • 8/3/2019 Phil. Trans. R. Soc. Lond. 2002, 357, 1481

    18/18

    1498 T. A. Moore and others Artificial antenna systems and reaction centres

    communication in multi-porphyrin light-harvesting arrays. J.

    Am. Chem. Soc. 116, 10 57810 592.

    Seth, J., Palaniappan, V., Wagner, R. W., Johnson, T. E.,

    Lindsey, J. S. & Bocian, D. F. 1996 Soluble multiporphyrin

    arrays. 3. Static spectroscopic and electrochemical probes of

    electronic communication. J. Am. Chem. Soc. 118,

    11 19411 207.

    Shreve, A. P., Trautman, J. K., Frank, H. A., Owens, T. G. &

    Albrecht, A. C. 1991 Femtosecond energy-transfer pro-

    cesses in the B800-850 light-harvesting complex of Rhodo-

    bacter sphaeroides 2.4.1. Biochim. Biophys. Acta 1058, 280

    288.

    Shreve, A. P., Trautman, J. K., Frank, H. A., Owens, T. G.,

    van Beek, J. B. & Albrecht, A. C. 1992 On subpicosecond

    excitation energy transfer in light harvesting complexes

    (LHC): the B800-B850 LHC of Rhodobacter sphaeroides

    2.4.1. J. Lumin. 53, 179186.

    Sieferman-Harms, D. 1987 The light-harvesting and protective

    functions of carotenoids in photosynthetic membranes.

    Physiol. Plant 69, 561568.

    Steinberg-Yfrach, G., Liddell, P. A., Hung, S.-C., Moore,

    A. L., Gust, D. & Moore, T. A. 1997 Conversion of light

    energy to proton potential in liposomes by artificial photo-

    synthetic reactions centres. Nature 385, 239241.

    Steinberg-Yfrach, G., Rigaud, J.-L., Durantini, E. N., Moore,

    A. L., Gust, D. & Moore, T. A. 1998 Light-driven pro-

    duction of ATP catalyzed by FoF1-ATP synthase in an arti-

    ficial photosynthetic membrane. Nature 392, 479482.

    Strachan, J. P., Gentemann, S., Seth, J., Kalsbeck, W. A.,

    Lindsey, J. S., Holten, D. & Bocian, D. F. 1997 Effects of

    orbital ordering on electronic communication in multipor-

    phyrin arrays. J. Am. Chem. Soc. 119, 11 19111 201.

    Trautman, J. K., Shreve, A. P., Owens, T. G. & Albrecht,

    A. C. 1990 Femtosecond dynamics of carotenoid-to-chloro-

    phyll energy transfer in thylakoid membrane preparations.

    Chem. Phys. Lett. 166, 369374.

    Truscott, T. G. 1990 The photophysics and photochemistry of

    the carotenoids. Photochem. Photobiol. B 6, 359371.

    Van Patten, P. G., Shreve, A. P., Lindsey, J. S. & Donohoe,

    R. J. 1998 Energy-transfer modeling for the rational design

    of multiporphyrin arrays. J. Phys. Chem. B 102, 42094216.

    Wagner, R. W. & Lindsey, J. S. 1994 A molecular photonic

    wire. J. Am. Chem. Soc. 116, 97599760.

    Wagner, R. W., Johnson, T. E. & Lindsey, J. S. 1996 a Soluble

    multiporphyrin arrays. 1. Modular design and synthesis. J.

    Am. Chem. Soc. 118, 11 16611 180.

    Wagner, R. W., Lindsey, J. S., Seth, J., Palaniappan, V. &

    Bocian, D. F. 1996b Molecular optoelectronic gates. J. Am.

    Chem. Soc. 118, 39963997.

    Wasielewski, M. R. 1992 Photoinduced electron transfer insupramolecular systems for artificial photosynthesis. Chem.

    Rev. 92, 435461.

    Wasielewski, M. R. & Kispert, L. D. 1986 Direct measurement

    of the lowest excited singlet state lifetime of all-trans--

    carotene and related carotenoids. Chem. Phys. Lett. 128,

    238243.

    Yamamoto, H. Y. 1979 Biochemistry of the violaxanthin cycle.

    Pure Appl. Chem. 51, 639648.

    Yamamoto, H. Y. & Bassi, R. 1995 Carotenoids: localization

    and function. In Oxygenic photocynthesis: the light reactions

    (ed. D. R. Ort & C. F. Yocum), pp. 539563. Dordrecht,

    The Netherlands: Kluwer.

    Phil Trans R Soc Lond B (2002)

    Discussion

    J. Barber (Department of Biological Sciences, Imperial Col-

    lege, London, UK). Why have you used C60 as an elec-

    tron acceptor?

    T. A. Moore. Other than being beautiful molecules,

    they can also act as multiple electron acceptors. More

    importantly, they have low reorganization energies and it

    was for this reason that the use of C60 allowed us to gener-ate recombination spin-polarized triplets similar to those

    observed in photosynthetic reaction centres.

    R. van Grondelle (Department of Biophysics, Free Univer-

    sity, Amsterdam, The Netherlands). In your artificial photo-

    synthetic systems, the light-harvesting function of

    carotenoids is emphasized. In nature, the photoprotective

    role of carotenoids seems to be the optimized function.

    Please comment.

    T. A. Moore. We have, indeed, in the past studied the

    protective effect of carotenoids, but in using them as a

    functional antenna, we automatically get the protective

    effect, bearing in mind that energy tr